• Aucun résultat trouvé

Comparing transient and steady-state methods for the thermal conductivity characterization of a borehole heat exchanger field in Bergen, Norway.

N/A
N/A
Protected

Academic year: 2021

Partager "Comparing transient and steady-state methods for the thermal conductivity characterization of a borehole heat exchanger field in Bergen, Norway."

Copied!
25
0
0

Texte intégral

(1)

1

Comparing transient and steady-state methods for the thermal

1

conductivity characterization of a borehole heat exchanger field in

2

Bergen, Norway

3

4

Nicolò Giordano 1,2, Jessica Chicco 2, Giuseppe Mandrone 2, Massimo Verdoya 3, Walter H. Wheeler 4 5

1 Institut national de la recherche scientifique – Centre Eau Terre Environnement, 490 Rue de la Couronne, G1K 9A9,

6

Québec (QC), Canada

7

2 Dipartimento di Scienze della Terra, Università di Torino, Via Valperga Caluso, 35 10125 Torino, Italy

8

3 Dipartimento di Scienze della Terra, dell’Ambiente e della Vita, Università di Genova, Viale Benedetto XV, 5-I 16132

9

Genova, Italy

10

4 NORCE Norwegian Research Centre, Energy department, Nygårdsgaten 112, 5008 Bergen, Norway

11 12

Corresponding author: 13

Nicolò Giordano – INRS, nicolo.giordano@ete.inrs.ca, ORCID 0000-0002-0747-444X 14

N.G. was formerly working at UNITO, to which the study is principally ascribable. 15

16

Key words: petrophysical properties, thermal conductivity, P-wave velocity, density, borehole heat 17 exchangers 18 19 Abstract 20

A comparative study was carried out aiming at characterizing the thermal conductivity of rocks 21

sampled in a borehole heat exchanger field. Twenty-three samples were analysed with four 22

different methods based on both steady-state and transient approaches: transient divided bar 23

(TDB), transient line source (TLS), optical scanning (OS) and guarded hot plate (GHP). Moreover, 24

mineral composition (from XRD analyses), P-wave velocity and density were investigated to assess 25

the petro-physical heterogeneity and to investigate possible causes of divergence between the 26

methods. The results of thermal conductivity showed that TLS systematically underestimates 27

thermal conductivity on rock samples by 10-30% compared to the other devices. The differences 28

between TDB and OS, and GHP and OS are smaller (about 6% and 10%, respectively). The average 29

deviation between TDB and GHP, for which the specimen preparation and the measurement 30

procedure was similar, is about 10%. In general, the differences are ascribable to sample 31

preparation, heterogeneity and anisotropy of the rocks, and contact thermal resistance, rather than 32

the intrinsic accuracy of the device. In case of good-quality and homogeneous samples, uncertainty 33

can be as low as 5%, but due to the above mentioned factors usually uncertainty is as large as 10%. 34

Opposite relationships between thermal conductivity and P-wave velocity were observed when 35

(2)

2

analysing parallel and perpendicular to the main rock foliation. Perpendicular conductivity values 36

grow with increasing perpendicular sonic velocity, while parallel values exhibit an inverse trend. 37

Thermal conductivity also appears to be inversely correlated to density. In quartz-rich samples, high 38

thermal conductivity and low density were observed. In samples with calcite or other likely dense 39

mineral phases, we noticed that lower thermal conductivity corresponds to higher density. The 40

presence of micas is likely to mask major differences between silicate and carbonate samples. 41

42

1. INTRODUCTION

43

The knowledge of the heat transfer processes in the underground is of utmost importance in several 44

fundamental geoscientific topics not only of general interest (e.g. tectonics, basin analysis, etc.), but 45

also in applied research with special reference to geothermal energy. Regarding geothermal 46

applications, in the last 30 years, ground source heat pumps (GSHP) systems have been developing 47

and widely spreading in the framework of the heating and cooling (H&C) of the residential, office 48

and commercial buildings. Major numbers are in North America, Europe and China (Lund and Boyd 49

2015). In 2014, direct uses of geothermal energy in Norway counted an installed capacity of around 50

1300 MWt, an annual energy consumption of 2300 GWh and a load factor of 0.2 (Lund and Boyd 51

2015). These are mostly related to GSHP, which have increased in number since 2000 (Midtømme 52

et al. 2015), and underground thermal energy storage (UTES; Cabeza 2015). In GSHP and UTES 53

applications, the Earth is exploited for direct use of the heat at accessible depths ( 100 m) by means 54

of either closed-loop borehole heat exchangers (BHE) or open-loop well doublets (e.g. Yang et al. 55

2010), and as a storage medium for sensible heat applications (e.g. Giordano et al. 2016 and 56

references therein). The performances of BHEs vary significantly depending on the rock type, but 57

also on the presence of groundwater flow which may enhance the heat transfer. 58

Precise and accurate thermal conductivity measurements of unconsolidated sediments and rocks 59

are crucial for a reliable definition of the heat transfer mechanism within geologic media. Thermal 60

conductivity and specific heat represent the most important properties to describe the mechanism 61

of heat transfer in any material. Studies of the thermo-physical properties of rocks mainly address 62

thermal conductivity because its range of variation is wider than specific heat. Conductivity is 63

primarily controlled by the mineral composition and the texture of the rock. It is generally an 64

anisotropic property, but for many rocks, the effects of anisotropy are minor compared to variations 65

in mineral composition. The bulk value of thermal conductivity generally increases with increasing 66

(3)

3

water saturation and density, and decreases with increasing porosity and temperature (Čermák and 67

Rybach 1982; Clauser and Huenges 1995; Alishaev et al. 2012; Mielke et al. 2017). However, 68

correlations between thermal and other properties are not always well defined, mainly due to 69

mineralogical, physical and geochemical factors (Kukkonen and Peltoniemi 1998). 70

In this paper, we investigate the thermal properties of rocks collected in the area of 200-m-deep 71

BHEs situated south of Bergen, Norway (60.34°N, 5.34°E; Fig. 1), which has been covering for the 20 72

years the H&C needs of a school building (Giordano et al. 2017). In the last years, the heat pumps 73

coefficient of performance has significantly decreased. As a part of the study was therefore 74

necessary to understand the causes of the thermal depletion and to evaluate the sustainable use of 75

the geothermal resource. As a first step of the analysis, here we present results of thermal 76

conductivity measurements of the lithotypes hosting the BHE field. 77

Over the past 40 years, the necessity of accurate data on both shallow and deep geothermal 78

reservoirs stimulated the development of new effective approaches and equipment for the 79

assessment of thermal conductivity (Pasquale et al. 2015; Popov et al. 2016). In this perspective, 80

comparative studies are important to evaluate the reliability and accuracy of the different methods 81

(e.g. Popov et al. 1999; Zhao et al. 2016) as well as comparisons among the several mixing laws 82

proposed in literature (e.g. Fuchs et al. 2013 and references therein). In this study, we used four 83

different measurement techniques and compared the results in order to evaluate the effects of 84

experimental conditions (e.g. sample preparation, measurement procedure, minimization of 85

thermal contact resistance) and to understand the potential error sources related to the various 86

techniques. In addition, mineral composition, compressional wave velocity and density were 87

detected to investigate the influence of the petro-physical heterogeneity on thermal conductivity 88

and possible causes of divergence in the results of the different techniques adopted. 89

90

2. MATERIALS AND METHODS

91

2.1 Samples and Geological Setting

92

This study focuses on 23 samples collected in close (100-500 m) proximity in a Silurian-aged thrust 93

complex (Fig. 1) on the outskirts of Bergen, Norway. The samples were collected from metric layers 94

analogous to those encountered in the geothermal boreholes. The study area belongs to the Minor 95

Bergen Arc tectonic unit (Proterozoic to Silurian), which together with Øygarden Complex 96

(Proterozoic), Ulriken Gneiss Complex (Proterozoic), Anorthosite Complex (Proterozoic) and the 97

(4)

4

Major Bergen Arc (Cambrian to Silurian) form the Bergen Arc System (Kolderup and Kolderup, 1940; 98

Fossen, 1989; Fossen and Ragnhildstveit, 2008; Fig. 1). The Minor Bergen Arc is a strongly deformed 99

continental basement-cover couplet that has been subdivided in the Nordåsvatn Complex, 100

Storetveit Group and Gamlehaugen Complex. 101

102

103

Figure 1 The geological setting of the study area (modified from Fossen and Ragnhildstveit 2008). OY – 104

Øygarden Complex, MBA – Minor Bergen Arc, GMH – Gamlehaugen Complex, NDV – Nordåsvatn Complex. 105

Coordinate system WGS84/UTM Zone 32N. 106

107

The first is a heterogeneous complex of meta-sedimentary (mafic micaschists with garnet and 108

amphibole in concentrations up to 25%) and meta-igneous (mainly fine-grained and coarse-grained 109

amphibolites and gabbros; serpentinites in smaller amounts) rocks interpreted as a strongly 110

dismembered ophiolite complex (Fossen 1989 and references therein). An unconformity separates 111

this complex to the supposed younger Storeveit Group, which occurs as a thin (less than 100 m) 112

continuous zone within the aforementioned unit and it is made of green metaconglomerates (with 113

(5)

5

clasts of trondhjemites, amphibolites and epidosites) of the Paradis Formation associated with 114

marbles and calcareous garnet-amphibolite micaschists of the Marmorøyen Formation (Fossen 115

1989 and references therein). Finally, the Gamlehaugen Complex groups strongly deformed 116

orthogneisses (proto- to ultra-mylonitic augen gneisses) and metasediments (quartz-schists with 117

abundant presence of mica and feldspar) interpreted as a detached continental basement-cover 118

sequence tectonically emplaced into the Nordåsvatn Complex (Fossen 1989 and references 119

therein). 120

121

Table 1 List of the collected samples. Acronyms: QMS (quartz-mica-schist); AM (amphibolite); AMS 122

(amphibolite-mica-schist); MB (marble); MS (mica-schist); AG (augen gneiss); QS (quartz-schist). 123

Number Lithology Latitude (N) Longitude (E) Description Geological Unit

1 QMS 60.347722 5.354263 Quartz-Mica-schist Nordåsvatn Complex 2 AM 60.345898 5.355659 Amphibolite 3 AM 60.345360 5.351639 Amphibolite 4 AM 60.345000 5.349383 Amphibolite 5 AMS 60.342757 5.346535 Amphibole-mica-schist 6 AM 60.343270 5.346782 Amphibolite 7 MB 60.344262 5.348224 Marble Storeveit Group 8 MB 60.344262 5.348224 Marble 9 MB 60.345252 5.347805 Marble 10 MB 60.345617 5.347565 Marble 11 MC 60.342172 5.347595 Meta-conglomerate 12 MS 60.342473 5.347279 Mica-schist Nordåsvatn Complex 13 MS 60.343494 5.345621 Mica-schist 14 MS 60.343512 5.341679 Mica-schist 15 QMS 60.344187 5.340576 Quartz-mica-schist Gamlehaugen Complex 16 AG 60.344180 5.339734 Augen gneiss 17 MS 60.347821 5.337358 Mica-schist 18 MS 60.347821 5.337358 Mica-schist 19 MS 60.347821 5.337358 Mica-schist 20 AG 60.341814 5.337834 Augen gneiss 21 QS 60.340960 5.335609 Quartz-schist 22 QS 60.340960 5.335609 Quartz-schist 23 QS 60.340960 5.335609 Quartz-schist 124

Samples locations were selected according to the surface and depth distribution of the different 125

geological units and to the role of a specific lithotype unit played in the underground thermal and 126

hydraulic regime (Tab. 1). Nine samples belong to the Nordåsvatn Complex (the most representative 127

(6)

6

unit), nine to the Gamlehaugen Complex (expected to be the group of rocks with the greatest 128

thermal conductivity) and five specimens were collected in the Storetveit Group (the marble 129

formation expected to play a crucial role in the fluid circulation). Specimens 7-10 of the Storetveit 130

Group are marbles of the Marmorøyen Formation, which crops out in the study area (even if this 131

detail is not shown in Fig. 1, as outcrops are very small and difficult to map). 132

133

2.2 Thermal Conductivity Measurement Techniques

134

Four specific devices with different basic principles were used in order to compare the thermal 135

conductivity datasets and discuss the divergence in terms of fundamental theory behind each of 136

them, sample preparation and rock heterogeneity. Measurements were carried out by means of the 137

transient divided bar (TDB) (Pasquale et al. 2015), the transient line source (TLS) (Bristow et al. 138

1994), the optical scanning (OS) (Popov et al. 2016) and the guarded hot plate (GHP) (Filla 1997). 139

These methods are briefly described in the following together with the specific devices used to 140

perform the measurements. 141

2.2.1 Transient Divided Bar (TDB) 142

The device includes a stack of elements consisting of two copper blocks of known thermal capacity, 143

between which the studied rock specimen (cylindrical shape) is interposed. The upper block has the 144

same diameter of the specimen and acts as a heat source; the lower block is larger and acts as a 145

heat sink. When the room temperature T0 is attained, the lower copper block is plunged into a

146

thermostatic bath with temperature 10 to 15 °C lower than T0. The heat flowing through the sample

147

is equal to the heat adsorbed by the sink and the thermal conductivity can be found by monitoring 148

the temperature changes of the source and the sink (Pasquale et al. 2015). From Fourier’s postulate, 149

the amount of heat removed from the upper copper block Cu ΔTu in a given period of time ∆𝑡 =

150 𝑡2− 𝑡1 is: 151 152 𝐶𝑢

∙ ∆𝑇

𝑢

=

𝜆𝑆 ℎ

∫ (𝑇

𝑢

− 𝑇

𝑙

) 𝑑𝑡

𝑡2 𝑡1

(1) 153 154

where Cu is the thermal capacity (J K-1) of the upper block at constant pressure, ΔTu (°C) the

155

temperature change of the upper block during a time step dt, λ (W m-1 K-1) the thermal conductivity

156

of the rock specimen, h (m) and S (m2) are the height and the cross-sectional area of the rock sample,

(7)

7

respectively, T is temperature and the suffixes u and l refer to upper and lower block. The change in 158

temperature is recorded by means of thermocouples connected to a digital acquisition system. 159

During the measurements, the temperature Tu and the difference Tu – Tl are recorded for a period

160

of at least 300 s. Thermal conductivity determinations are carried out at time steps of 10 s and 161

generally about 10 different time steps are analysed for each sample to obtain an average value. 162

Thermal conductivity can be obtained from: 163

164

𝜆 =

𝐶𝑢∙∆𝑇𝑢∙ℎ

𝑆∙(𝑇̅̅̅̅−𝑇𝑢 ̅̅̅)∙∆𝑡𝑙

(2)

165

where 𝑇̅̅̅ − 𝑇𝑢 ̅𝑙 is the average difference between the values of Tu and Tl during the time step Δt.

166

Two correction factors are also taken into account. The first is related to the heat coming from the 167

specimen and therefore an effective heat capacity must be considered, that is the sum of Cu and

168

one third of the rock heat capacity (measured by means of a water calorimeter). The second 169

depends on the heat transfer from the surrounding environment to the upper block and can be 170

easily estimated by operating under steady state conditions, i.e. 2 hours after the beginning of the 171

test (Pasquale et al. 2015). The tests were carried out in a temperature-controlled environment at 172

20 °C and the thermal conductivity was determined with an accuracy of ± 5%. 173

2.2.2 Transient Line Source (TLS) 174

The line source method is based on the generation of heat at a constant rate by a heated wire. If 175

the line source is assumed to be infinitely long and infinitely thin, fully immersed in an infinite and 176

homogeneous medium, and a recording thermistor is placed in the same probe, the temperature 177

response, in a constant room temperature environment, can be described by: 178 179 𝑇2

− 𝑇

1

= −

𝑞′ 4𝜋𝜆

∙ ln(

𝑡2 𝑡2−𝑡1

) (3)

180 181

where q’ (W m-1) is the specific rate at which the heat is generated and T

1 and T2 (°C) are the

182

recorded temperatures at time steps t1 and t2 (s) respectively. This approximation of the general

183

heat transfer equation of an infinite line source allows for thermal conductivity measurements 184

with errors within ± 10% if the early time data (t < 200 s) are ignored and only the straight line in 185

a graph ΔT vs ln t is considered. 186

(8)

8

The apparatus adopted in this study is the commercial K2DPro Thermal Property Analyzer 187

(Decagon Devices) that fully complies with the standards ASTM D5334 and IEEE 442 (ASTM, 2014; 188

IEEE, 2003). A proprietary algorithm fits time and temperature data with exponential integral 189

functions using a non-linear least squares method. The single needle sensor RK-1 (length 6 cm, 190

diameter 3.9 mm), specific for hard rock samples, was used. A thin hole (4.5 mm in diameter) was 191

drilled into the sample in order to host the probe and alumina thermal grease (9 W m-1 K-1) was

192

applied to the probe in order to minimize the sensor/rock contact resistance. The measurements 193

were carried out in a controlled room temperature environment and the heating power was about 194

6.5 ± 0.5 W m-1 for all the whole set of samples. A 600 s measuring time was adopted (maximum

195

available) so that 300 s of heating and 300 s of cooling were recorded with a 10 s sampling interval. 196

At least 3 measurements were carried out on each sample and an average was taken. A time 197

interval of 15 min for each test was adopted in order to allow for the thermal re-equilibrium 198

between the sample and the probe. Before each triplet of measurements, a calibration was 199

performed with the verification standard provided and the calculated calibration factor was 200

applied to correct the thermal conductivity values (ASTM 2014). 201

2.2.3 Optical Scanning (OS) 202

The Optical Scanning is a precision non-contact method designed by Popov et al. (1999). It was 203

developed since the 80s by means of several comparisons with different techniques and calibrated 204

with a number of standard materials (Popov et al. 2016 and references therein). The measurement 205

procedure consists of scanning the sample of a surface with a focused movable heat source (electric 206

lamp) in combination with three infrared temperature sensors (Popov et al. 1999). The heat source 207

and sensors move with the same speed (controllable from 2 to 10 mm s-1) relative to the sample

208

and at a constant spacing xo among each other (adjustable from 20 to 100 mm). Rock samples are

209

placed on a platform in order to be scanned from below by the trolley containing the optical source 210

and the sensors. A synthetic black enamel is applied on the surface in order to neglect the influence 211

of optically transparent surfaces or different minerals’ reflectance. The technical parameters of the 212

apparatus adopted in this study are the ones described in Popov et al. 2016 for the Type 1. 213

The method is based on the heat conduction equation for a quasi-stationary field in a movable 214

coordinate system. The temperature rise induced in the sample and recorded along the scanning 215

line is related to the thermal conductivity as 216

𝑇

− 𝑇

𝑐

=

𝑞

2𝜋∙𝑥0∙𝜆 (4)

(9)

9

where Th and Tc (°C) are temperatures registered by the hot (after the heating source) and cold

218

(before) sensors, q (W) is the constant heating power and xo is the distance between the source and

219

the hot sensor. Since the temperature rise depends on both the heating power and the value of xo,

220

the measurements are always performed in comparison to standards of known thermal conductivity 221

(e.g. glasses, fused quartz, gabbro etc.). Standards are placed before and after the sample under 222

exam such that they are scanned with the same q and x0 and within the same room temperature.

223

In this study, an OS apparatus designed by Lippman and Rauen GbR was used. For all the 224

measurements q was set to about 17 W (20% of the maximum power) with a scanning velocity of 5 225

mm s-1 and x

0 = 50 mm. The standards adopted were homogeneous gabbro samples (provided by

226

the company) with thermal conductivity of 2.37 W m-1 K-1 and the measurements were performed

227

in a controlled temperature environment at around 20 °C. Three runs were carried out per each 228

sample’s scanning line and an average was taken. The accuracy certified by the company is ± 3%. In 229

case of a sample with a clear two-dimensional anisotropy (e.g. schists), the principal values of the 230

conductivity were determined from two non-collinear scanning lines performed on one face not 231

parallel to the foliation as suggested by Popov et al. 2016. This was done for all the samples in order 232

to check a possible thermal anisotropy even where a clear textural anisotropy was not present (e.g. 233

marbles). An anisotropy factor KT = λpar/λper was then calculated and the sample was considered

234

thermally homogeneous only in case of KT < 1.1.

235

2.2.4 Guarded Hot Plate (GHP) 236

The guarded hot plate or heat flow meter is a stationary technique based on standards ASTM C177 237

(2013) and ASTM C518 (2017) that also allows measuring thermal conductivity of semiconductors 238

at high temperatures (Filla 1997). A steady state one-dimensional heat flow is applied through a 239

specimen by two parallel plates guarded at different constant temperatures, while the whole stack 240

is insulated to avoid side heat losses. Temperature and heat flow are continuously registered on 241

both plates throughout the test by means of thermocouples and transducers, while an axial load is 242

provided to minimize the thermal contact resistance. 243

The apparatus adopted in this study is the commercial device Fox50 (LaserComp Inc., 2001-2004). 244

It can measure the thermal conductivity of cylindrical shaped samples, with diameters of 245

25 ÷ 61 mm and maximum thickness of 25 mm, in the range -5 ÷ 185 °C (Raymond et al. 2017). 246

Proprietary heat flow transducers together with high accuracy (± 0.01 °C) type E thermocouples are 247

bonded and sealed to the surfaces of both plates. Guarded temperature on the heat source (upper 248

(10)

10

plate) and sink (lower) are provided by Peltier elements and a downward heat flow is generated 249

through the rock specimen. From the Fourier heat conduction law, the temperature gradient within 250

the sample is given by the difference between the cold and hot plate temperatures divided by the 251

sample thickness. However, for conductivity values > 0.2-0.3 W m-1 K-1 (basically all the rocks and

252

minerals), the temperatures on the sample surfaces are different from the plates because the 253

thermal contact resistance is not significantly smaller than the sample thermal resistance. 254

Therefore, the temperature difference between the upper and lower plates is given by: 255

∆𝑇

𝑝𝑙𝑎𝑡𝑒𝑠

= 𝛿𝑇

𝑢

+ ∆𝑇

𝑠𝑎𝑚𝑝𝑙𝑒

+ 𝛿𝑇

𝑙 (5) 256

where δTu and δTl (°C) are the temperature differences between upper plate and sample surface,

257

and between sample surface and lower plate, respectively. The contact thermal resistance R (m2 K

258 W-1) is: 259

𝑅 =

𝛿𝑇 𝑞′′

(6) 260

where q’’ (W m-2) is the heat flow recorded through each plate. R depends on the type of material,

261

the interface pressure applied and the roughness of the sample. The electric signal Q (V) recorded 262

by the heat flow transducers is proportional to the heat flow q’’ through a calibration factor 263

Scal [W m-2 V-1] that is determined using standard materials with known thermal conductivity. Q,

264

which is recorded during the experiment, is related to the thermal conductivity as: 265

𝑄 =

𝑞′′ 𝑆𝑐𝑎𝑙

=

∆𝑇𝑝𝑙𝑎𝑡𝑒𝑠 (∆𝑥 𝜆+2𝑅)∙𝑆𝑐𝑎𝑙 (7) 266

where Δx is the thickness of the sample. The absolute accuracy of the device is ± 3% in the 267

conductivity range of 0.1 ÷ 10 W m-1 K-1. Silicon or glycerine paste or rubber pads of known thermal

268

resistance can be employed to smooth the problem of contact resistance. The measurements were 269

carried out at 20 °C with a ΔT = 10 °C between the upper (25 °C) and lower (15 °C) plate and the 270

average of 10 sets of measurements was taken as final value. 271

272

2.3 Sample preparation

273

The whole sample collection was divided into two main datasets: dataset1 and dataset2. Dataset1 274

included all the 23 rock samples while dataset2 represents a subset of dataset1, namely nine 275

representative samples. Regarding thermal conductivity (see Section 2.2), most of samples of 276

(11)

11

dataset1 were analysed with two methods (OS and TLS), whereas dataset2 was additionally tested 277

with other two techniques (TDB and GHP). 278

Thermal conductivity in all 23 samples was studied in the two main directions, i.e. parallel (λpar) and

279

perpendicular (λper) to the main foliation. Five of the samples showed no clear foliation and KT < 1.1;

280

these are classified as thermally homogeneous and an average effective value was reported. This 281

procedure was adopted in both OS and TLS. For the measurements with OS, the samples were cut 282

according to the foliation in order to obtain two perpendicular polished surfaces upon which the 283

coating layer was applied. For the TLS analyses, previously cut and polished sample surfaces were 284

drilled with a 4 mm rotary hammer bit to host the RK-1 single needle sensor. Two perpendicular 285

drillings were performed in eleven out of twenty-three samples. Parallel and perpendicular 286

conductivity values were calculated with the same methodology adopted for OS (Popov et al. 2016) 287

and the anisotropy factor calculated. 288

Owing to issues related to obtain samples with the size and characteristics required by both TDB 289

and GHP, cylindrical rock specimens were prepared from samples of dataset2 by means of a 290

diamond-head corer and, through the use of a fine abrasive, both surfaces were rubbed down to 291

get flat (within 0.1 mm), parallel and smooth surfaces (within 0.03 mm). Finally, the obtained core 292

specimens had cylindrical shape, 25 ± 0.1 mm in diameter and 20 ± 0.5 mm in thickness. To improve 293

the contact between rock specimen and the blocks (TDB) and plates (GHP), a film of silicone paste 294

of about 0.1 mm was smeared on both samples surfaces. 295

Samples for OS and TLS are relatively easy to prepare, requiring a cut and painted surface (OS) or a 296

drilled hole (TLS). In contrast TDB and GHP require core drilling followed by precision grinding, 297

making their preparation critical to the success of the analysis. 298

299

2.4 Compressional wave velocity and density

300

The compressional wave velocity test is a non-destructive method based on the principle that pulse 301

velocity of ultrasound waves, propagating through a solid material, depends on the density and the 302

elastic properties of that material (Al-Khafaji and Purnell 2016). The commercial device PUNDIT Lab 303

(by Proceq Switzerland) was used in this study. The apparatus is coupled with a pair of transducers 304

that transmit and receive waves with a central frequency of 54 kHz, working on ASTM D2845 (2008) 305

procedure. The tests were performed on the samples as analysed by the OS technique, as flat 306

surfaces are necessary to make good contact between the specimen and the transducers. On each 307

(12)

12

sample, 25-30 measurements were carried out to get accurate results and the average between 308

these acquisitions was calculated. 309

Density was obtained by weighing the samples after water saturation under vacuum and measuring 310

their volume through immersion in water. Samples were then oven-dried at 70 °C for one day in 311

order to obtain the dry mass and to infer porosity. All the samples denoted negligible porosity (< 3%) 312

and thus water content was assumed to be of negligible importance on the measurements of petro-313 physical properties. 314 315 2.5 Mineralogical composition (XRD) 316

The mineralogical composition of the samples of dataset2 was investigated by means of powder 317

X-Ray diffraction (XRD) on crystalline samples. The measurements were performed with the Philips 318

X’Pert-Pro device (Malvern Panalytical ©2018) consisting of a Bragg-Brentano geometry and 319

equipped with a stationary, centrally placed, X-ray tube. The tube was operated using a CuKα 320

radiation at 40 mA, 40 kV and 1.5417 Å. Spectra were recorded in the 2θ range 5-70° with a 15 s 321

counting time and 0.008° 2θ step. 322

A qualitative analysis was firstly performed, distinguishing between two defined phase structures 323

of calcite and quartz present in each sample. The peak shape was modelled with a Pseudo-Voigt 324

function of which, the FWHM (Full Width of Half Maximum), was refined as a function of 2θ taking 325

into account both Gaussian and Lorentzian broadening. The refinement was carried out in 326

particular in the space group R-3c (calcite), P3221 (quartz).

327

In order to get an alternative estimate of the accuracy of the refined structural data, a comparison 328

among the set of structural parameters obtained using different refinement strategies on the same 329

diffraction data was carried out. These comparisons show that realistic estimates of the error bars 330

are ± 0.001 Å for the cell parameters. The error in the estimation of the phase content is ± 1% wt. 331 332 3. RESULTS 333 3.1 Dataset1 334

The bulk thermal conductivity values of dataset1 were measured by means of OS in the University 335

of Bergen’s laboratory (Tab. 2) and with TLS in the University of Torino’s laboratory (Tab. 3). The 336

samples coded B (e.g. 1_B) indicate that the specific rock specimens were measured in Bergen only; 337

those coded T (e.g. 1_T) indicate that they were analysed in Torino only; when the specimen is called 338

(13)

13

BT (e.g. 4_BT) means that the same rock specimen was measured by both techniques. This should 339

warn the reader that differences between 1_B and 1_T can even be related to the sample 340

heterogeneity and not only to the adopted techniques. Some samples (3_T, 9_BT, 13_BT, 17_BT, 341

18_BT, 19_BT) broke during preparation and it was not possible to analyse neither thermal 342

conductivity with TLS or OS, nor sonic velocity. 343

The standard deviations reported for OS are related to the values measured along the scanning lines 344

and thus due to the intrinsic heterogeneity of the rock samples. These cannot be compared to the 345

standard deviations of the TLS, which relate to the repeatability and precision of the KD2 Pro. The 346

standard deviations of the effective thermal conductivity (three runs along each scanning line) 347

measured with the OS were ± 0.7% on average, with a maximum of ± 1.6% and a minimum of ± 348

0.1%. OS and TLS bulk thermal conductivity was calculated as a geometric average between parallel 349

and perpendicular values. KT refers to the anisotropy factor given by the parallel to perpendicular

350

values ratio. 351

Mica-schists and amphibolites of the Nordåsvatn Complex show significant thermal anisotropy, with 352

parallel values higher than perpendicular ones by 50% in OS and 46% in TLS data. The marbles of 353

the Storeveit Group present an isotropic nature as expected, with anisotropy factor always smaller 354

than 1.1. The Gamlehaugen Complex shows clear thermal anisotropy in the quartz-schists and 355

quartz-mica-schists (KT > 1.2), in contrast to the general isotropic texture of both augen and

356

mylonitic gneisses (KT < 1.1). In absolute terms, the whole dataset is rather homogeneous, a bit

357

surprising given the presence of high quartz-content lithotypes. Low values of quartz-schists 21, 22 358

and 23 can be explained by abundant presence of muscovite, which presents strong anisotropy in 359

thermal conductivity (0.6 W m-1 K-1 perpendicular, 3.9 W m-1 K-1 parallel; Clauser and Huenges

360

1995). The OS average thermal conductivity is 2.75 W m-1 K-1 with a standard deviation of 0.29; TLS

361

data records an average conductivity of 2.18 W m-1 K-1 with a minimum of 1.37 and maximum of

362

3.16 (standard deviation 0.51). 363

364

Table 2 Thermal conductivity λ (W m-1 K-1) and anisotropy factor K

T of dataset1 measured by means of OS.

365

Sample ID Lithology λbulk λperp St. dev. λpar St. dev. KT Comments

1_B QMS 3.088 2.159 0.675 4.416 1.168 2.05 2_B AM 2.681 2.322 0.484 3.095 0.273 1.33 3_B AM 2.606 2.137 0.724 3.178 0.239 1.49 4_BT AM 3.128 2.652 0.508 3.689 0.351 1.39 5_BT AMS 2.682 2.373 0.436 3.031 0.239 1.28 6_BT AM 2.643 2.165 0.415 3.226 0.335 1.49

(14)

14 7_BT MB 2.934 / / 2.934 0.174 / isotropic 8_BT MB 2.854 / / 2.854 0.128 / isotropic 9_BT MB 2.778 2.671 0.415 2.889 0.125 1.08 10_BT MB 2.995 2.913 0.621 3.079 0.194 1.06 11_BT MC 2.328 1.888 0.867 2.870 0.552 1.52 12_BT MS 2.826 2.461 0.470 3.246 0.199 1.32 13_BT MS 2.332 1.620 0.976 3.356 1.447 2.07 14_B MS 2.423 2.228 0.391 2.636 0.290 1.18 15_B QMS 3.333 3.233 0.540 3.437 0.169 1.06 16_B AG 2.955 / / 2.955 0.189 / isotropic 17_BT MS 2.275 2.069 0.269 2.500 0.379 1.21 18_BT MS 2.906 / / 2.906 0.862 / isotropic 20_B AG 2.946 2.924 0.645 2.969 0.299 1.02 21_B QS 2.677 2.286 1.288 3.134 0.563 1.37 22_BT QS 2.842 2.300 0.884 3.511 0.308 1.53 23_B QS 2.301 2.103 0.425 2.518 0.280 1.20 366

Table 3 Thermal conductivity λ (W m-1 K-1) values and anisotropy factor K

T obtained with TLS.

367

Sample ID Lithology λbulk λperp St. dev. λpar St. dev. KT Comments

1_T QMS 1.367 1.006 0.060 1.859 0.074 1.85

2_T AM 2.422 2.745 0.072 2.137 0.104 0.78

4_BT * AM 2.743 / / 2.743 0.097 / only perp. drilling possible

5_BT * AMS 2.266 / / 2.266 0.045 / only perp. drilling possible

6_BT AM 1.911 1.613 0.035 2.263 0.107 1.40

7_BT MB 2.518 / / 2.518 0.127 / isotropic

8_BT MB 2.448 / / 2.448 0.175 / isotropic

10_BT ** MB 2.317 2.317 0.137 / / / only par. drilling possible

11_BT MC 1.797 1.291 0.096 2.501 0.165 1.94 12_BT MS 2.042 1.538 0.130 2.712 0.055 1.76 14_T MS 1.773 1.454 0.084 2.162 0.067 1.49 15_T QMS 3.023 2.513 0.184 3.636 0.095 1.45 16_T AG 3.155 / / 3.156 0.456 / isotropic 20_T AG 2.066 1.856 0.073 2.299 0.146 1.24

21_T * QS 1.426 / / 1.426 0.034 / only perp. drilling possible

22_BT QS 2.175 1.779 0.088 2.659 0.152 1.50

23_T QS 1.621 1.699 0.038 1.547 0.042 0.91

* parallel value; ** perpendicular value 368

369

Comparing the results of the different techniques, we observe a general underestimation of TLS 370

with respect to OS, both for the bulk conductivity and λpar and λper (Fig. 2). TLS underestimates

371

thermal conductivity with respect to OS, with a maximum difference of 56% (sample 1) in the bulk 372

value. Generally, the higher biases occur in the parallel analyses (58%) than in those perpendicular 373

(15)

15

(54%). By taking OS as a reference, 68% of the TLS results underestimate the thermal conductivity 374

of the rock samples analysed by a minimum of 10% to a maximum of 30%; for the 24%, the 375

underestimate is more than 30% and only few outliers (8%) overestimate it. A significant difference 376

among bulk, parallel and perpendicular values is not clearly evident in Figure 2. If only the BT 377

samples are taken into consideration, the underestimate of TLS with respect to OS is on average 378

20.1%, 20.3% and 27.5% for bulk, parallel and perpendicular values respectively, with a minimum 379

of 12.3% (4_BT bulk) and maximum of 37.5% (12_BT perpendicular). 380

381

382

Figure 2 Comparison between OS and TLS bulk, parallel and perpendicular values of thermal conductivity 383

384

No clear direct correlation between bulk thermal conductivity (from OS technique) and P-wave 385

velocity is observed in dataset1 (Fig. 3). Mielke et al. (2017) also showed that low-porosity rocks, as 386

those investigated in this paper, present weak correlations. It is nevertheless interesting to note the 387

contrast between properties when considered parallel and perpendicular to the main rock foliation. 388

Perpendicular conductivity values grow with increasing perpendicular sonic velocity, while parallel 389

values exhibit an inverse trend. Therefore, it is shown that only along the direction perpendicular to 390

foliation heat and P-waves propagation in the medium follow similar patterns. Conversely, when 391

analysing the sample along the main foliation, propagation paths might not be necessarily the same. 392

Unfortunately, due to the heterogeneity of the collection and the limited number of samples, this 393

cannot be related to specific rock features. It would be necessary to investigate further, but this 394

goes beyond the purpose of the present paper. When comparing the anisotropy factors (Tab. 4), 395

(16)

16

the data are quite similar, with eight out of twelve samples showing a KS/KT ratio within the 0.85 ÷

396

1.15 range (i.e. 15% bias) and the best results occurring in the BT samples (same specimen analysed 397

by both techniques). 398

399

400

Figure 3 Relationship between thermal conductivity (OS technique) and P-wave velocity. 401

402

Table 4 P-wave velocity Vp (m s-1), sonic (K

S) and thermal anisotropy (KT from OS) of dataset1. N is the

403

number of measurements. 404

Sample ID Lithology Vpbulk Vpperp St.

dev. N Vppar St. dev. N KS KT KS / KT 1_T / 1_B MS 3448 2691 45 27 4418 353 28 1.64 2.05 0.80 2_T / 2_B AM 4769 3953 154 27 5753 165 27 1.46 1.33 1.09 4_BT AM 3727 2978 114 27 4664 136 27 1.57 1.39 1.13 5_BT AMS 4805 4038 124 28 5719 171 27 1.42 1.28 1.11 6_BT AM 4836 3284 92 28 7123 296 27 2.17 1.49 1.46 10_BT MB 5877 5774 274 26 5982 107 27 1.04 1.06 0.98 11_BT MC 3082 2435 27 26 3901 118 26 1.60 1.52 1.05 12_BT MS 3295 2963 73 26 3664 58 26 1.24 1.32 0.94 14_T / 14 B MS 4804 4209 127 26 5483 226 27 1.30 1.18 1.10 21_T / 21_B QS 2533 1187 57 27 5407 209 27 4.55 1.37 3.32 22_BT QS 3099 2343 37 27 4100 123 27 1.75 1.53 1.15 23_T / 23_B QS 4301 3577 128 27 5172 81 27 1.45 1.20 1.21 405 3.2 Dataset2 406

(17)

17

Dataset2 comprised nine of the dataset1 samples further studied using GHP and TLS. Of the nine 407

specimens, 3 are perpendicular and one is parallel to the main foliation; the rest are from isotropic 408

samples. Cylindrical shape specimens were obtained from the samples analysed with the OS and 409

TLS techniques (samples BT) or by the TLS only (samples T). 410

411

Table 5 Comparison of thermal conductivity results of dataset2 measured with the four techniques (see 412

text); the accuracy of each technique is in brackets. Last column reports average values from literature 413

(aČermák and Rybach 1982; bKukkonen and Peltoniemi 1998; cDi Sipio et al. 2014; dEppelbaum et al. 2014;

414

eRamstadt et al. 2015; fMielke et al. 2017).

415

Sample ID Lithology Orientation TDB

(±5%) OS (±5%) GHP (±3%) TLS (±10%) Literature 1_T QMS Parallel 4.23 4.42* 3.98 1.86 3.8d 7_BT MB Isotropic 2.56 2.93 2.76 2.52 2.7 ÷ 2.8 a,f 10_BT MB Isotropic 2.90 3.00 2.62 2.32** 2.7 ÷ 2.8 a,f 11_BT MC Isotropic 2.29 2.33 2.91 1.80 2.6 e 12_BT MS Perpendicular 2.89 2.46 2.32 1.54 2.5 ÷ 2.8 d,e 14_T MS Parallel 2.44 2.64* 2.38 2.16 2.5 ÷ 2.8 d,e 15_T QMS Parallel 3.36 3.44* 3.96 3.64 3.8 d 16_T AG Isotropic 2.91 2.96* 2.88 3.16 2.4 ÷ 3.0 a,b,c,e 20_T AG Isotropic 2.83 2.95* 2.90 2.07 2.4 ÷ 3.0 a,b,c,e

* the B specimen was measured; ** perpendicular value 416

417

The results are consistent with average values in literature (Tab. 5). The analysis of dataset2 418

confirms that TLS underestimates thermal conductivity with respect to TDB (-26% on average) and 419

GHP (-24% on average), with a maximum of more than 50% in sample 1_T, which is highly 420

anisotropic. It is worth stressing that in four samples (7_BT, 14_T, 15_T and 16_T) the bias with GHP 421

is within the 10% accuracy expected for the TLS device. In particular, in the very homogeneous and 422

isotropic augen gneiss 16_T, TLS registers a thermal conductivity higher than TDB, OS and GHP by 423

9%, 7% and 10% respectively. A conductivity greater than TDB (8%) and OS (6%) is also given by TLS 424

in the same lithology of 15_T (Fig. 4). 425

On the contrary, a good general agreement between OS, TDB and GHP is observed (Fig. 5): a 7-8% 426

divergence from the overall mean is registered compared to an 18% average bias recorded by TLS. 427

In the five samples in which the TLS bias is greater than the accuracy of the devices used for the 428

analyses (error bars do not overlap), that value was discarded and a new average calculated (Tab. 429

6). The new values were then adopted to compare the different techniques. TDB and OS show an

430

average divergence of 6% with maximum of 16.8% and 13.7% in 12_BT and 7_BT respectively. 431

(18)

18

Moreover, TDB and OS are within 5% in six out of nine samples. Comparisons between GHP and OS, 432

and GHP and TDB have worse accordance, with average 9.8% and 10.4% respectively, and only in 433

two and three cases better than 5%. The greatest biases are observed in BT samples, wherein the 434

same specimen was analysed by means of the three techniques: in particular, the largest deviations 435

(24.7% and 22.3%) were obtained between TDB and GHP, which used the same core sample. The 436

discrepancy could be attributed to the difference between static and transient analyses: in three 437

out of nine samples (11_BT, 12_BT and 15_T), TDB and GHP error bars do not overlap. 438

439

440

Figure 4 Thermal conductivity results of dataset2. 441

442 443 444

Table 6 Comparison of thermal conductivity results for dataset2. 445 Sample ID Lithology Mean all methods (W m-1 K-1)

Differences to mean Recalculated

mean w/o TLS (W m-1 K-1)

Differences between pairs TDB [%] OS [%] GHP [%] TLS [%] TDB-GHP [%] TDB-OS [%] OS-GHP [%] 1_T QMS 3.62 17.0 21.7 9.8 -48.6 4.20 6.2 4.2 10.2 7_BT MB 2.69 -4.9 8.8 2.5 -6.4 2.69* 7.4 13.7 6.3 10_BT MB 2.71 7.1 10.5 -3.3 -14.3 2.84 9.9 3.2 13.1 11_BT MC 2.33 -1.8 -0.1 24.8 -22.8 2.51 24.7 1.6 23.1 12_BT MS 2.30 25.5 6.8 0.8 -33.1 2.56 22.3 16.8 5.5 14_T MS 2.41 1.5 9.8 -1.0 -10.2 2.41* 2.5 8.3 10.8 15_T QMS 3.60 -6.6 -4.4 9.9 1.2 3.60* 16.5 2.2 14.3 16_T AG 2.98 -2.2 -0.6 -3.4 6.2 2.98* 1.1 1.7 2.8 20_T AG 2.69 5.3 9.8 7.9 -23.0 2.89 2.4 4.2 1.7

(19)

19 * value of the first column

446 447

448

Figure 5 Comparison of conductivity measurement techniques for dataset2.

449 450

451

Figure 6 XRD results of the main mineral groups of dataset2 samples. 452

453

XRD analyses show the bulk mineralogy of the samples, entirely quartz and clay/mica except two 454

with carbonate (Fig. 6). These analyses were performed because many of the samples have a 455

microcrystalline structure, so the recognition of the mineral phases with optical techniques is often 456

difficult and in many cases not useful from a quantitative point of view. Quartz and micas are the 457

main mineral phases of the mica-schists and gneisses, while calcite is predominant in the marbles 458

(20)

20

(samples 10_BT and 7_BT; Fig. 6). As expected, Table 6 shows that higher quartz content 459

corresponds to higher thermal conductivity. For example, the quartz-mica schist samples 1_T and 460

15_T have mean conductivities of 4.2 and 3.6 W m-1 K-1 respectively. These samples also have low

461

density, consistent with the low relative density of quartz and mica (Fig. 7). The high content of 462

micas (likely muscovite and biotite) might mask significant differences between silicate and 463

carbonate samples, given that calcite thermal conductivity (3.2÷3.6 W m-1 K-1; Clauser and Huenges

464

1995; Andolfsson 2013) is higher than micas (2.0÷2.3 W m-1 K-1; Clauser and Huenges 1995). In

465

samples with calcite or likely other dense mineral phases (e.g. olivine), not clearly detected in the 466

XRD study, we noticed lower thermal conductivity which is accompanied by higher density values 467

(e.g. 10_BT and 11_BT, Fig. 7). The thermal conductivity and density of 10_BT is consistent with 468

isotropic marble. 11_BT is characterized by the highest density and lowest conductivity. Its density 469

and thermal conductivity are consistent with the facies description of clasts of trondjemite, 470

epidosite and amphibolite (Fossen 1989). Density ranges for these rocks are: trondjemite 2.7, 471

epidosite 2.8÷3.0 and amphibolite 2.9÷3.1 g cm-3; thermal conductivity: trondjemite 1.8÷2.6,

472

epidosite 2.4÷4.5 and amphibolite 2.2÷2.9 W m-1 K-1 (Popov et al. 1999; Miao et al. 2014, Merriman

473

et al. 2013). The XRD shows relatively low peaks for quartz and mica, and, at slightly lower 2-theta 474

than quartz, some additional low peaks not present in the other samples (Fig. 6). 475

476

Figure 7 Comparison between thermal conductivity and density in dataset2. 477

478

4. DISCUSSION

479

Among the techniques used for the thermal conductivity analyses of our samples, it is reported in 480

the literature that TLS is a valid method to measure highly porous, soft materials (e.g. Di Sipio et al. 481

2014), but with hard materials the results may often be affected by large biases. In our analyses, 482

particular care has been adopted in sample preparation. Each sample was previously prepared with 483

(21)

21

two parallel surfaces in order to have a 90° ± 5° angle with the drill bit; a drill press was used to drill 484

the sample to assure a precise perpendicular drilling; water was constantly poured in to safeguard 485

the widia-diamond bit and facilitate the exit of drill cutting; the hole was cleaned up by removing 486

dust and cuttings with compressed air after the end of drilling; the probe and the inner hole were 487

carefully coated with thermal grease; finally, the probe was inserted slowly to allow trapped air to 488

escape. 489

Despite the careful sample preparation, the TLS systematically showed low values with respect to 490

the other three techniques. TLS gave lower conductivity than OS (more than 90% of samples in 491

dataset1) and GHP and TDB (more than 65% in dataset2). Anisotropy seems to have some influence, 492

the average bias being 14% in isotropic against 27% in anisotropic rocks, even if isotropic samples 493

10_BT and 20_T showed biases of 23 and 30%. By increasing the TLS values by 20%, more than half 494

of the samples would be within 10% of the OS value (dataset1). However, in dataset2 this would 495

cause severe overestimation for samples 7_BT, 15_T and 16_T. In summary, we infer the “high” 496

values are correct, and attribute the “low” values to problems with deficiencies in the sample 497

preparation for TLS measurements. Clearly, TLS strongly depends on sample preparation and 498

heterogeneity and anisotropy of the rock. Isotropic, homogeneous and competent (weathered 499

samples could break during preparation) lithotypes can limit the bias to < 10% if care is taken in the 500

measurement procedure above explained. For anisotropic and highly heterogeneous rocks, and 501

weak or weathered samples, the other methods are preferable to avoid significant underestimation 502

of the actual thermal conductivity. 503

Even though the GHP and the TDB methods are similar in sample preparation and precautions to 504

limit the contact resistance were adopted (in two cases, 11_BT and 12_BT, the specimens were even 505

the same), the results were similar in samples 7_BT, 16_T and 20_T, but significantly different in 506

samples 10_BT, 11_BT, 12_BT and 15_T. A systematic explanation for the variation in results is not 507

evident. Only minor influence on the results was found for the heterogeneity and anisotropy of the 508

samples. It is nevertheless evident that the core preparation is crucial to get reliable results; but it 509

also seems that other factors are important, such as the quality of the measuring stack (contact 510

between plates/blocks and sample, accuracy of temperature sensors/heat flow meters) and the 511

employment of standards (TDB) rather than previously obtained calibration curves (GHP). 512

Smaller biases were obtained between GHP and OS, and TDB and OS, with the TDB-OS differences 513

being consistently smaller than the others. Differences with OS are also related to the different 514

measuring procedures. The OS scans the entire sample while the other techniques give instead more 515

(22)

22

localized information. Thus, the OS can thoroughly characterize a rock sample with information 516

about heterogeneity and thermal anisotropy in an easier way in comparison to the other three 517

techniques. Another advantage of the OS is the simple sample preparation. Last but not least, the 518

problem of thermal contact resistance is bypassed. On the other hand, samples for OS analysis must 519

be big enough to neglect boundary effects (Popov et al. 2016), even if the energy input (heat and 520

speed) can be varied accordingly. 521

Finally, it is important to stress that OS is the fastest way to analyse thermal conductivity, not only 522

in terms of single measurement (1-2 min against 5-8 min with TDB and 60-70 min with GHP), but 523

also for the simplest sample preparation. The OS seems to be the best with friable and weathered 524

samples since there is no contact and only a flat surface is required, but micro-cracks are still a 525

problem. Nevertheless, if the sample is competent enough, drilling several cores with different 526

orientations allows evaluating thermal anisotropy also with GHP or TDB. 527

Bulk thermal conductivity and sonic velocity values did not show a clear direct correlation in the 528

collections under study, herein attributed to the low porosity of the samples (Mielke et al. 2017). 529

This could be related to the opposite trends observed in the parallel and perpendicular analyses, 530

that show an inverse and a direct relationship respectively. In the authors’ knowledge, this has not 531

been reported in literature so far and deserves to be investigated further. 532

A clearly defined correlation between thermal conductivity and density was not recorded. However, 533

in quartz-rich samples we observed high thermal conductivity and low density as also observed by 534

Pasquale et al. (2015). On the contrary, in calcite-rich samples we noticed lower thermal 535

conductivity not related to a well-defined density value. The presence of micas is likely to mask 536

major differences between silicate and carbonate samples. 537

538

5. CONCLUSIONS

539

A comparison among four different laboratory methodologies to analyse the rock thermal 540

conductivity was carried out. Steady state (GHP) and transient (TLS, TDB, OS) methods were adopted 541

and results compared to highlight qualities and flaws of the different techniques. Moreover, 542

compressional wave velocities, density and mineral composition were investigated. The results of 543

this study are preparatory for future activities that will encompass the set-up of numerical modelling 544

of the underground thermal structure of the BHE field in Bergen. 545

Among the four methods for measuring thermal conductivity, even if steady state techniques are 546

expected to be more accurate, our results indicate that TDB and OS give more congruent results. 547

(23)

23

TLS, instead, systematically underestimates thermal conductivity in the investigated samples, 548

confirming that it is hardly applicable to hard rocks. Due to heterogeneity, anisotropy and 549

mechanical properties of the rocks, the use of at least two different techniques seems 550

recommendable in investigations on rock thermal properties. An uncertainty of 5 to 10% is the best 551

that one can expect even in good-quality and homogeneous samples. Geothermal modelling often 552

relies on values of thermal conductivity without well-defined uncertainty boundaries. The inclusion 553

of this uncertainty may increase the reliability of estimations. 554

555

Acknowledgments

556

The authors would like to thank the laboratory of the University of Bergen, in particular Niels Bo 557

Jensen, for carrying out measurements and processing the data with the TCS. Giorgia Confalonieri 558

and Alessandro Pavese, are also warmly thanked for the XRD analyses. The authors are extremely 559

grateful to Jasmin Raymond for the use of the GHP in the Geothermal Open Lab at the Institut 560

national de la recherche scientifique of Québec, Canada. 561

562

Author Contributions

563

N.G. collected the samples and took care of the samples preparation for OS and TLS, measurements 564

and data processing of thermal conductivity with TLS; he also performed P-wave analyses and 565

processed the data. J.C. prepared the samples, carried out analyses and processed the data of TDB 566

and GHP; she also took care of XRD processing and density analyses. N.G. and J.C. wrote together 567

the original draft paper, finalizing figures and tables. W.H.W. and G.M. conceptualized the original 568

idea of the study, and together with M.V. advised on the rigorous experimental analyses and revised 569 the manuscript. 570 571 References 572 573

Alishaev MG, Abdulagatov IM, Abdulagatova ZZ (2012) Effective thermal conductivity of fluid-574

saturated rocks. Experiment and modeling. Eng Geol 135-136:24-39. 575

Al-Khafaji S, Purnell P (2016) Effect of coupling media on ultrasonic pulse velocity in concrete: a 576

preliminary investigation. Int J Civ Env Eng 10:118-121. 577

Andolfsson T (2013) Analyses of thermal conductivity from mineral composition and analyses by use 578

of Thermal Conductivity Scanner: A study of thermal properties in Scanian rock types. 579

Dissertation, Lund University. 580

(24)

24

ASTM C177 (2013) Standard test method for steady-state heat flux measurements and thermal 581

transmission properties by means of the guarded-hot-plate apparatus. ASTM International, 582

West Conshohocken, PA, doi: 10.1520/C0177-13. 583

ASTM D5334 (2014) Standard test method for determination of thermal conductivity of soil and soft 584

rock by thermal needle probe procedure. ASTM International, West Conshohocken, PA, doi: 585

10.1520/D5334-14. 586

ASTM C518 (2017) Standard test method for steady-state thermal transmission properties by means 587

of the heat flow meter apparatus. ASTM International, West Conshohocken, PA, doi: 588

10.1520/C0518-17. 589

ASTM D2845 (2008) Standard test method for laboratory determination of pulse velocities and 590

ultrasonic elastic constants of rock. ASTM International, West Conshohocken, PA, doi: 591

10.1520/D2845-08. 592

Bristow KL, White RD, Kluitenberg GJ (1994) Comparison of single and dual probes for measuring 593

soil thermal properties with transient heating. Austr J Soil Res 32:447-464. 594

Cabeza LF (2015) Advances in thermal energy storage systems – methods and applications. 595

Woodhead Publishing, Cambridge, UK. 596

Čermák VL, Rybach L (1982) Thermal conductivity and specific heat of minerals and rocks. In: 597

Angenheister G (ed) Numerical Data and Functional Relationships in Science and Technology, 598

New Series, Group V (Geophysics and Space Research), Vol. 1a (Physical Properties of Rocks), 599

Springer, Berlin, pp 305-343. 600

Clauser C, Huenges E (1995) Thermal conductivity of rocks and minerals. Rock physics & phase 601

relations: a handbook of physical constants, American Geophysical Union, pp 105-126. 602

Di Sipio E, Galgaro A, Destro E, Teza G, Chiesa S, Giaretta A, Manzella A (2014) Subsurface thermal 603

conductivity assessment in Calabria (southern Italy): a regional case study. Env Earth Sci 604

72:1383-1401. 605

Eppelbaum L, Kutasov I, Pilchin A (2014) Thermal properties of rocks and density of fluids. In: Applied 606

Geothermics. Springer, Berlin, pp 99-149. 607

Filla BJ (1997) A steady-state high-temperature apparatus for measuring thermal conductivity of 608

ceramics. Rev Sci Ins 68(7):2822-2829. 609

Fossen H (1989) Geology of the Minor Bergen Arc, West Norway. Norg Geol Unders Bull, 416:47-62. 610

Fossen H, Ragnhildstveit J (2008) Berggrunnskart 1115 I, Scale 1:50.000, Norge Geol Unders. 611

Fuchs S, Schütz F, Förster HJ, Förster A (2013) Evaluation of common mixing models for calculating 612

bulk thermal conductivity of sedimentary rocks: correction charts and new conversion 613

equations. Geoth 47:40-52. 614

Giordano N, Chicco J, Bastesen E, Wheeler WH, Mandrone G (2017) Thermo-hydraulic 615

characterization of a fractured shallow reservoir in Bergen (Norway) to improve the efficiency 616

of a BHE field. European Geosciences Union, General Assembly, April 23-28, Wien, Austria, 617

EGU2017-18581. 618

(25)

25

Giordano N, Comina C, Mandrone G, Cagni A (2016) Borehole thermal energy storage (BTES). First 619

results from the injection phase of a living lab built up in unsaturated alluvial deposits (Torino, 620

IT). Ren En 86:993-1008. 621

IEEE Guide for Soil Thermal Resistivity Measurements (2003). IEEE Std 442-1981 1–16. 622

Kolderup CF, Kolderup NH (1940) Geology of the Bergen Arc System. Bergen Museums Skrifter, 20, 623

137 p. 624

Kukkonen IT, Peltoniemi S (1998) Relationships between thermal and other petrophysical 625

properties of rocks in Finland. Phys Chem Earth 23(3):341-349. 626

Lund JW, Boyd TL (2015) Direct utilization of geothermal energy 2015 worldwide review. 627

Proceedings World Geothermal Congress, n. 01000, April 19-25, Melbourne, Australia. 628

Merriman JD, Whittington AG, Hofmeister AM, Nabelek PI, Benn K (2013). Thermal transport 629

properties of major Archean rock types to high temperature and implications for cratonic 630

geotherms. Precambrian Res 233:358-372. 631

Miao SQ, Li HP, Chen G (2014). Temperature dependence of thermal diffusivity, specific heat 632

capacity, and thermal conductivity for several types of rocks. J Therm Anal Calorim 115(2): 1057-633

1063. 634

Midttømme K, Ramstad RK, Müller J (2015) Geothermal energy – Country update for Norway. 635

Proceedings World Geothermal Congress, n. 01071, April 19-25, Melbourne, Australia. 636

Mielke P, Bär K, Sass I (2017) Determining the relationship of thermal conductivity and 637

compressional wave velocity of common rock types as a basis for reservoir characterization. J 638

App Geophys 140:135-144. 639

Pasquale V, Verdoya M, Chiozzi P (2015) Measurements of rock thermal conductivity with a 640

Transient Divided Bar. Geoth, 53:183-189. 641

Popov YA, Beardsmore G, Clauser C, Roy S (2016) ISRM Suggested methods for determining thermal 642

properties of rocks from laboratory tests at atmospheric pressure. Rock Mechanics and Rock 643

Engineering 49:4179-4207. 644

Popov YA, Pribnow DFC, Sass JH, Williams CF, Burkhardt H (1999) Characterization of rock thermal 645

conductivity by high-resolution optical scanning. Geoth 28:253-276. 646

Ramstad RK, Midttømme K, Liebel TH, Frengstad BS, WIllemoes-Wissing B (2015) Thermal 647

conductivity map of the Oslo region based on thermal diffusivity measurements of rock core 648

samples. Bull Eng Geol Env 74:1275-1286. 649

Raymond J, Comeau F-A, Malo M, Blessent D, López Sánchez IJ (2017) The geothermal open 650

laboratory: a free space to measure thermal and hydraulic properties of geological materials. 651

IGCP636 Annual Meeting, Santiago de Chile, Chile, 1-3. 652

Yang H, Cui P, Fang Z, (2010) Vertical-borehole ground-coupled heat pumps: a review of models and 653

systems. App En 87:16-27. 654

Zhao D, Quian X, Gu X, Ayub Jajjia S, Yang R (2016) Measurement techniques for thermal 655

conductivity and interfacial thermal conductance of bulk and thin film materials. J Elect 656

Pack 138(4):040802. 657

Figure

Figure 1 The geological setting of the study area (modified from Fossen and Ragnhildstveit 2008)
Table  1  List  of  the  collected  samples.  Acronyms:  QMS  (quartz-mica-schist);  AM  (amphibolite);  AMS 122
Table 2 Thermal conductivity λ (W m -1  K -1 ) and anisotropy factor K
Table 3 Thermal conductivity λ (W m -1  K -1 ) values and anisotropy factor K
+7

Références

Documents relatifs